Update to v0.2:

Changed: more precise statement of the bound on the large x decay of u (Theorem 1.2)

  Added: more details in the proof of the large x decay of u (Theorem 1.2)

  Fixed: miscellaneous typos.
This commit is contained in:
Ian Jauslin 2021-05-22 18:38:30 -04:00
parent f8a94b06fa
commit ed63f14f57
2 changed files with 93 additions and 23 deletions

View File

@ -125,10 +125,10 @@ Along with the computation of the low density energy of the simple equation in o
for $N$ particles in a cubic box of finite volume $V$ with periodic boundary conditions. for $N$ particles in a cubic box of finite volume $V$ with periodic boundary conditions.
The ground state eigenfunction $\psi_0$ is unique and non-negative, as can be shown using the Perron-Frobenius theorem, and thus we may normalize $\psi_0$ to obtain a probability measure. This is not the usual probability measure associated to a quantum state, which would be quadratic in the wave function, but since $\psi_0$ is non-negative and integrable ($\|\psi_0\|_1 \leqslant V^{1/2}\|\psi_0\|_2$), we may use it directly to define a probability measure, and this is the starting point of \cite{Li63}. Because particles interact pairwise, the ground state energy and other observables can be calculated in terms of the two-point correlation function associated to this probability measure: The ground state eigenfunction $\psi_0$ is unique and non-negative, as can be shown using the Perron-Frobenius theorem, and thus we may normalize $\psi_0$ to obtain a probability measure. This is not the usual probability measure associated to a quantum state, which would be quadratic in the wave function, but since $\psi_0$ is non-negative and integrable ($\|\psi_0\|_1 \leqslant V^{1/2}\|\psi_0\|_2$), we may use it directly to define a probability measure, and this is the starting point of \cite{Li63}. Because particles interact pairwise, the ground state energy and other observables can be calculated in terms of the two-point correlation function associated to this probability measure:
\begin{equation} \begin{equation}
g_N(x_1-x_2) := \lim_{N,V\to\infty, N/V = \rho}\frac{V^2\int dx_3\cdots dx_N\ \psi_0(x_1,x_2,x_3,\dots,x_N)}{\int dy_1\cdots dy_N\ \psi_0(y_1,\dots,y_N )} g(x_1-x_2) := \lim_{N,V\to\infty, N/V = \rho}\frac{V^2\int dx_3\cdots dx_N\ \psi_0(x_1,x_2,x_3,\dots,x_N)}{\int dy_1\cdots dy_N\ \psi_0(y_1,\dots,y_N )}
\end{equation} \end{equation}
In \cite{Li63}, under a few physically motivated approximations, in the thermodynamic limit, in which the number of particles $N$ and the volume of the gas $V$ are taken to infinity, with $\rho:=\frac NV$ fixed, an equation for the limiting two-point correlation function $g_{\infty}$ is derived. In \cite{Li63}, under a few physically motivated approximations, in the thermodynamic limit, in which the number of particles $N$ and the volume of the gas $V$ are taken to infinity, with $\rho:=\frac NV$ fixed, an equation for the limiting two-point correlation function $g$ is derived.
The function $u(x)$ in \eqref{simp} is then defined as $u(x):=1- g_{\infty}(x)$. Note that since by definition $g_{\infty}(x) \geqslant 0$, $u(x) \leqslant 1$. The function $u(x)$ in \eqref{simp} is then defined as $u(x):=1- g(x)$. Note that since by definition $g(x) \geqslant 0$, $u(x) \leqslant 1$.
Because the expected values in the ground state of many physical observables can be calculated in terms of $g$, any method for computing $g$ that bypasses directly solving the $N$-body Schr\"odinger equation for the Hamiltonian \eqref{ham0} provides an effective means for the computation of these values, and this motivates the study of the simple equation system\-~(\ref{simp}). Indeed, the ground state energy per particle is given in terms of $g$ by the second equation in \eqref{simp}. There is so far no rigorous derivation of \eqref{simp} from the $N$-body Schr\"odinger equation, and hence there is no mathematical Because the expected values in the ground state of many physical observables can be calculated in terms of $g$, any method for computing $g$ that bypasses directly solving the $N$-body Schr\"odinger equation for the Hamiltonian \eqref{ham0} provides an effective means for the computation of these values, and this motivates the study of the simple equation system\-~(\ref{simp}). Indeed, the ground state energy per particle is given in terms of $g$ by the second equation in \eqref{simp}. There is so far no rigorous derivation of \eqref{simp} from the $N$-body Schr\"odinger equation, and hence there is no mathematical
@ -325,7 +325,7 @@ We shall also need various $L^p$ bounds on $u'$, and for these we need a detaile
If $(1+ |x|^4)v(x)\in L^1(\mathbb R^3)\cap L^2(\mathbb R^3)$, then If $(1+ |x|^4)v(x)\in L^1(\mathbb R^3)\cap L^2(\mathbb R^3)$, then
\begin{equation} \begin{equation}
\rho u(x)= \rho u(x)=
\frac{\sqrt{2+\beta}}{2\pi^2\sqrt{e}}\frac1{1+|x|^{4}} \frac{\sqrt{2+\beta}}{2\pi^2\sqrt{e}}\frac1{|x|^{4}}
+ +
R(x) R(x)
\label{udecay} \label{udecay}
@ -336,18 +336,13 @@ We shall also need various $L^p$ bounds on $u'$, and for these we need a detaile
, ,
\end{equation} \end{equation}
and where $|x|^4R(x)$ is in $L^2(\mathbb R^3)\cap L^\infty(\R^3)$, uniformly in $e$ on all compact sets. and where $|x|^4R(x)$ is in $L^2(\mathbb R^3)\cap L^\infty(\R^3)$, uniformly in $e$ on all compact sets.
Moreover, there is a constant $C$ independent of $e$ and $\rho$ such that for all $x$ Moreover, for every $\rho_0>0$, there is a constant $C$ that only depends on $\rho_0$ such that for all $x$, for all $\rho<\rho_0$,
\begin{equation}\label{fourdecay} \begin{equation}\label{fourdecay}
u(x) \leqslant C \rho^{-1}e^{-1/2}\frac{1}{1 + |x|^4}\ . u(x) \leqslant \min\left\{1,\frac C{\rho e^{\frac12}|x|^4}\right\}\ .
\end{equation} \end{equation}
\end{theorem} \end{theorem}
\begin{remark}
Theorem\-~\ref{theo:pointwise} actually holds with the condition that $|x|^4 v\in L^1(\mathbb R^3)$, without necessarily being in $L^2(\mathbb R^3)$, together with the condition $v\in L^1(\R^3)\cap L^2(\R^3)$, but the proof of that is a bit more involved, and will not be presented here.
\end{remark}
The next two theorems concern the monotonicity of $\rho\mapsto e(\rho)$ and convexity if $\rho\mapsto\rho e(\rho)$. The next two theorems concern the monotonicity of $\rho\mapsto e(\rho)$ and convexity if $\rho\mapsto\rho e(\rho)$.
These were conjectured in\-~\cite{CJL20}, and here, we prove them for small density $\rho$ (and, in the case of the monotonicity, also for large density). These were conjectured in\-~\cite{CJL20}, and here, we prove them for small density $\rho$ (and, in the case of the monotonicity, also for large density).
@ -600,7 +595,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
\begin{remark} \begin{remark}
As stated at the very beginning of the paper, we assume that $v$ is spherically symmetric. As stated at the very beginning of the paper, we assume that $v$ is spherically symmetric.
This is, however, used very little in the proofs. This is, however, used very little in the proofs.
In fact, the only theorem which relies on the spherical symmetry is theorem\-~\ref{theo:pointwise}. In fact, the only theorem that relies on the spherical symmetry is theorem\-~\ref{theo:pointwise}.
We believe it should still hold (provided the decay constant in\-~(\ref{udecay}) is suitably adapted) without the spherical symmetry. We believe it should still hold (provided the decay constant in\-~(\ref{udecay}) is suitably adapted) without the spherical symmetry.
In this case, the other theorems would not require the spherical symmetry. In this case, the other theorems would not require the spherical symmetry.
\end{remark} \end{remark}
@ -665,6 +660,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
U_1(x)=\frac{e^{\frac32}}\pi e^{-2\sqrt e|x|} U_1(x)=\frac{e^{\frac32}}\pi e^{-2\sqrt e|x|}
-\frac1\pi \left(\delta(x)+\frac{(\beta+1)e}\pi\frac{e^{-2\sqrt e|x|}}{|x|}\right) -\frac1\pi \left(\delta(x)+\frac{(\beta+1)e}\pi\frac{e^{-2\sqrt e|x|}}{|x|}\right)
\ast f_1\ast f_2 \ast f_1\ast f_2
\label{U1}
\end{equation} \end{equation}
where where
\begin{equation} \begin{equation}
@ -676,11 +672,13 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
=\frac{e}{\pi|x|} \int_0^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2} dt\ , =\frac{e}{\pi|x|} \int_0^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2} dt\ ,
\end{equation} \end{equation}
now, for all $T>0$, now, for all $T>0$,
\begin{eqnarray*} \begin{eqnarray}
\int_0^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}\ dt &=& \int_0^{T} e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}+\int_{T}^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}\ dt\\ \int_0^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}\ dt &=& \int_0^{T} e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}+\int_{T}^\infty e^{-\sqrt{2+\beta + t}(2\sqrt{e}|x|)} t^{-1/2}\ dt
&\leqslant& \int_0^{T} e^{-\sqrt{2+\beta}(2\sqrt{e}|x|)} t^{-1/2}+\int_{T}^\infty e^{-\sqrt{ t}(2\sqrt{e}|x|)} t^{-1/2}\ dt\\ \nonumber\\
&=& 2 T^{1/2} e^{-\sqrt{2+\beta}(2\sqrt{e}|x|)} + \frac{1}{\sqrt{e}|x|}e^{-\sqrt{T}(2\sqrt{e}|x|)}\ . &\leqslant&
\end{eqnarray*} 2 T^{1/2} e^{-\sqrt{2+\beta}(2\sqrt{e}|x|)} + \frac{1}{\sqrt{e}|x|}e^{-\sqrt{T}(2\sqrt{e}|x|)}\ .
\label{boundf2}
\end{eqnarray}
Choosing $T = 2+\beta$, we see that for large $(2\sqrt{e}|x|)$, $0 \leqslant f_2(x) \leqslant C e^{-\sqrt{2+\beta}(2\sqrt{e}|x|)}$. Choosing $T = 2+\beta$, we see that for large $(2\sqrt{e}|x|)$, $0 \leqslant f_2(x) \leqslant C e^{-\sqrt{2+\beta}(2\sqrt{e}|x|)}$.
Furthermore, Furthermore,
\begin{equation} \begin{equation}
@ -689,14 +687,16 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
\frac{e^{\frac32}}{\pi^3}\frac1{|x|^2}\ast g \frac{e^{\frac32}}{\pi^3}\frac1{|x|^2}\ast g
,\quad ,\quad
g(x)=\frac{(1-\sqrt e|x|)e^{-(2\sqrt e)|x|}}{|x|} g(x)=\frac{(1-\sqrt e|x|)e^{-(2\sqrt e)|x|}}{|x|}
\label{g}
\end{equation} \end{equation}
Using Using
\begin{equation} \begin{equation}
\frac{1}{|x-y|^2} = \frac{1}{1+|x|^2} + \frac{1-|y|^2 + 2x\cdot y}{ (1+|x|^2)|x-y|^2} \frac{1}{|x-y|^2} = \frac{1}{|x|^2} + \frac{-|y|^2 + 2x\cdot y}{|x|^2|x-y|^2}
\end{equation} \end{equation}
twice and the fact that $g(y)$ is even, integrates to zero, and $\int y g(y)\ dy=0$, twice and the fact that $g(y)$ is even, integrates to zero, and $\int y g(y)\ dy=0$,
\begin{equation} \begin{equation}
f_1(x) = \frac{1}{(1+|x|^2)^2} \frac{e^{\frac32}}{\pi^3}\left( - \int_{\R^3} |y|^2 g(y){\rm d} y + \int_{\R^3} \frac{(1-|y|^2 + 2x\cdot y)^2}{ |x-y|^2}g(y) {\rm d} y\right) f_1(x) = \frac{1}{|x|^4} \frac{e^{\frac32}}{\pi^3}\left( - \int_{\R^3} |y|^2 g(y){\rm d} y + \int_{\R^3} \frac{(-|y|^2 + 2x\cdot y)^2}{|x-y|^2}g(y) {\rm d} y\right)
\label{f1}
\end{equation} \end{equation}
We compute We compute
$ $
@ -708,7 +708,58 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
Therefore, Therefore,
\begin{equation} \begin{equation}
\lim_{|x|\to \infty} |x|^4 f_1(x) = -\frac{1}{2\pi^2\sqrt e} \qquad{\rm and}\qquad \lim_{|x|\to \infty} |x|^4 U_1(x) = \frac{1}{2\pi^2\sqrt e}\sqrt{2+\beta}\ . \lim_{|x|\to \infty} |x|^4 f_1(x) = -\frac{1}{2\pi^2\sqrt e} \qquad{\rm and}\qquad \lim_{|x|\to \infty} |x|^4 U_1(x) = \frac{1}{2\pi^2\sqrt e}\sqrt{2+\beta}\ .
\label{U1decay}
\end{equation} \end{equation}
\bigskip
\indent
We now turn to an upper bound of $U_1$.
First of all, if $|x|\leqslant\frac1{\sqrt e}$, then by\-~(\ref{g}) and\-~(\ref{f1}),
\begin{equation}
f_1(x)\geqslant 0
\end{equation}
and if $|x|>\frac1{\sqrt e}$, then
\begin{equation}
f_1(x)\geqslant
-\frac{1}{|x|^4} \frac{e^{2}}{\pi^3}\int_{\mathbb R^3} \frac{(-|y|^2 + 2x\cdot y)^2}{ |x-y|^2}e^{-(2\sqrt e)|y|} {\rm d} y
.
\end{equation}
We split the integral into two parts: $|y-x|>|x|$ and $|y-x|<|x|$.
We have, (recalling $|x|>\frac1{\sqrt e}$),
\begin{equation}
\int_{|y-x|>|x|} \frac{(-|y|^2 + 2x\cdot y)^2}{ |x-y|^2}e^{-(2\sqrt e)|y|} {\rm d} y
\leqslant
e^{-\frac52}C
\end{equation}
for some constant $C$ (we use a notation where the constant $C$ may change from one line to the next).
Now,
\begin{equation}
\int_{|y-x|<|x|} \frac{(-|y|^2 + 2x\cdot y)^2}{ |x-y|^2}e^{-(2\sqrt e)|y|} {\rm d} y
\leqslant
e^{-\sqrt e|x|}\int_{|y-x|<|x|} \frac{(|y|^2 + 2|x||y|)^2}{ |x-y|^2} {\rm d} y
\leqslant
|x|^5e^{-\sqrt e|x|}C
.
\end{equation}
Therefore, for all $x$,
\begin{equation}
f_1(x)\geqslant
-\frac{1}{|x|^4}C(e^{-\frac12}+e^2|x|^4e^{-\sqrt e|x|})
.
\end{equation}
Finally, by use\-~(\ref{boundf2}),
\begin{equation}
|x|^4\left(\delta(x)+\frac{(\beta+1)e}{\pi}\frac{e^{-2\sqrt e|x|}}{|x|}\right)\ast f_1\ast f_2(x)\geqslant
-Ce^{-\frac12}
.
\end{equation}
All in all, by\-~(\ref{U1}), (since $|x|^4e^{\frac32}e^{-2\sqrt e|x|}<Ce^{-\frac12}$)
\begin{equation}
|x|^4U_1(x)\leqslant Ce^{-\frac12}
.
\label{boundU1}
\end{equation}
\bigskip
\point \point
We now show that $\Delta^2\widehat U_2$ is integrable and square-integrable. We now show that $\Delta^2\widehat U_2$ is integrable and square-integrable.
@ -716,6 +767,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
\begin{equation} \begin{equation}
16e^2\Delta^2\equiv\partial_\kappa^4+\frac4\kappa\partial_\kappa^3 16e^2\Delta^2\equiv\partial_\kappa^4+\frac4\kappa\partial_\kappa^3
. .
\label{D2}
\end{equation} \end{equation}
We have, by the Leibniz rule, We have, by the Leibniz rule,
\begin{equation} \begin{equation}
@ -739,6 +791,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
for some family of constants $c_{l_1,\cdots,l_p}^{(p,n)}$ which can easily be computed explicitly, but this is not needed. for some family of constants $c_{l_1,\cdots,l_p}^{(p,n)}$ which can easily be computed explicitly, but this is not needed.
Now, since $S\geqslant 0$, $\frac\rho{1e}|\widehat S|\leqslant 1$, so $|\zeta_1|\leqslant\frac12$ and $\zeta_1=\frac12$ if and only if $\kappa=0$. Now, since $S\geqslant 0$, $\frac\rho{1e}|\widehat S|\leqslant 1$, so $|\zeta_1|\leqslant\frac12$ and $\zeta_1=\frac12$ if and only if $\kappa=0$.
Therefore, $\widehat U_2$ is bounded when $\kappa$ is away from 0, so it suffices to show that $\Delta^2\widehat U_2$ is integrable and square integrable at infinity and at 0. Therefore, $\widehat U_2$ is bounded when $\kappa$ is away from 0, so it suffices to show that $\Delta^2\widehat U_2$ is integrable and square integrable at infinity and at 0.
\bigskip
\subpoint We first consider the behavior at infinity, and assume that $\kappa$ is sufficiently large. \subpoint We first consider the behavior at infinity, and assume that $\kappa$ is sufficiently large.
@ -746,7 +799,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
To prove the corresponding claim for $\zeta_1$, we use the fact that $|x|^4 v$ square integrable, which implies that $\widehat S$ is as well. To prove the corresponding claim for $\zeta_1$, we use the fact that $|x|^4 v$ square integrable, which implies that $\widehat S$ is as well.
Therefore, by\-~(\ref{zetas}) for $0\leqslant n\leqslant 4$, $\kappa^2\partial_\kappa^n\zeta_1$ is integrable at infinity, and, therefore, square-integrable at infinity. Therefore, by\-~(\ref{zetas}) for $0\leqslant n\leqslant 4$, $\kappa^2\partial_\kappa^n\zeta_1$ is integrable at infinity, and, therefore, square-integrable at infinity.
Furthermore, by\-~(\ref{zetas}), $\zeta_1<\frac12-\epsilon$ for large $\kappa$, and $\partial^n\zeta_1$ is bounded, so $\partial_\kappa^{n-i}(\kappa^2+1)\partial_\kappa^i(1-\sqrt{1-2\zeta_1})$ is integrable and square integrable. Furthermore, by\-~(\ref{zetas}), $\zeta_1<\frac12-\epsilon$ for large $\kappa$, and $\partial^n\zeta_1$ is bounded, so $\partial_\kappa^{n-i}(\kappa^2+1)\partial_\kappa^i(1-\sqrt{1-2\zeta_1})$ is integrable and square integrable.
\bigskip
\subpoint As $\kappa\to0$ \subpoint As $\kappa\to0$
\begin{equation} \begin{equation}
@ -787,7 +840,7 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
=O(\kappa^{1-i}) =O(\kappa^{1-i})
. .
\end{equation} \end{equation}
Thus, by\-~(\ref{leibnitz}), Thus, by\-~(\ref{leibnitz}), as $\kappa\to0$,
\begin{equation} \begin{equation}
|\partial_\kappa^4\widehat U_2|=O(\kappa^{-1}) |\partial_\kappa^4\widehat U_2|=O(\kappa^{-1})
,\quad ,\quad
@ -795,7 +848,15 @@ This provides the control on $\|u'\|_q$ that we need to prove the theorem on co
. .
\end{equation} \end{equation}
Thus, $\Delta^2\widehat U_2$ is integrable and square integrable. Thus, $\Delta^2\widehat U_2$ is integrable and square integrable.
And since the $O(\cdot)$ hold uniformly in $e$ on all compact sets, $U_2|x|^4$ is bounded and square integrable uniformly in $e$ on all compact sets. And since the $O(\cdot)$ hold uniformly in $e$ on all compact sets, by\-~(\ref{D2}),
\begin{equation}
|x|^4U_2(x)
\leqslant
\frac{8e^{\frac32}}{16e^2}\int \left(\partial_{|k|}^4+\frac4{|k|}\partial_{|k|}^3\right)\hat U_2(|k|)\ dk
\leqslant\frac C{\sqrt e}
.
\end{equation}
This along with\-~(\ref{U1decay}) and\-~(\ref{boundU1}) implies\-~(\ref{udecay}) and\-~(\ref{fourdecay}).
\qed \qed
@ -1849,7 +1910,7 @@ $$
$$ $$
\end{proof} \end{proof}
It now follows that with $e_\star$ define as in Theorem~\ref{Mon}, on account of the bound on $\rho'$ proved there, and on account of Theorem~\ref{theo:pointwise} that there is a constant independent of $e$ such that for all $e\leqslant e_\star$, It now follows that with $e_\star$ defined as in Theorem~\ref{Mon}, on account of the bound on $\rho'$ proved there, and on account of Theorem~\ref{theo:pointwise} that there is a constant independent of $e$ such that for all $e\leqslant e_\star$,
\begin{equation}\label{varphipointwise} \begin{equation}\label{varphipointwise}
|\varphi(x)| \leqslant Ce^{-3/2}(1 + |x|^4)^{-1}\ . |\varphi(x)| \leqslant Ce^{-3/2}(1 + |x|^4)^{-1}\ .
\end{equation} \end{equation}

View File

@ -1,3 +1,12 @@
v0.2:
* Changed: more precise statement of the bound on the large x decay of u (Theorem 1.2)
* Added: more details in the proof of the large x decay of u (Theorem 1.2)
* Fixed: miscellaneous typos.
v0.1: v0.1:
* Added: references to [Nelson, 1964] and [Reed, Simon, 1975]. * Added: references to [Nelson, 1964] and [Reed, Simon, 1975].